Valorization of Bambara groundnut shell via intermediate ... Valorization of...5 and Feroz Kabir...

25
Valorization of Bambara groundnut shell via intermediate pyrolysis: Products distribution and characterization MOHAMMED, Isah Yakub, ABAKR, Yousif Abdalla, MUSA, Mukhtar, YUSUP, Suzana, SINGH, Ajit and KABIR, Feroz <http://orcid.org/0000-0002-3121- 9086> Available from Sheffield Hallam University Research Archive (SHURA) at: http://shura.shu.ac.uk/13402/ This document is the author deposited version. You are advised to consult the publisher's version if you wish to cite from it. Published version MOHAMMED, Isah Yakub, ABAKR, Yousif Abdalla, MUSA, Mukhtar, YUSUP, Suzana, SINGH, Ajit and KABIR, Feroz (2016). Valorization of Bambara groundnut shell via intermediate pyrolysis: Products distribution and characterization. Journal of Cleaner Production, 139, 717-728. Copyright and re-use policy See http://shura.shu.ac.uk/information.html Sheffield Hallam University Research Archive http://shura.shu.ac.uk

Transcript of Valorization of Bambara groundnut shell via intermediate ... Valorization of...5 and Feroz Kabir...

  • Valorization of Bambara groundnut shell via intermediate pyrolysis: Products distribution and characterizationMOHAMMED, Isah Yakub, ABAKR, Yousif Abdalla, MUSA, Mukhtar, YUSUP, Suzana, SINGH, Ajit and KABIR, Feroz

    Available from Sheffield Hallam University Research Archive (SHURA) at:

    http://shura.shu.ac.uk/13402/

    This document is the author deposited version. You are advised to consult the publisher's version if you wish to cite from it.

    Published version

    MOHAMMED, Isah Yakub, ABAKR, Yousif Abdalla, MUSA, Mukhtar, YUSUP, Suzana, SINGH, Ajit and KABIR, Feroz (2016). Valorization of Bambara groundnut shell via intermediate pyrolysis: Products distribution and characterization. Journal of Cleaner Production, 139, 717-728.

    Copyright and re-use policy

    See http://shura.shu.ac.uk/information.html

    Sheffield Hallam University Research Archivehttp://shura.shu.ac.uk

    http://shura.shu.ac.uk/http://shura.shu.ac.uk/information.html

  • Page 1 of 24

    Valorization of Bambara Groundnut Shell via Intermediate Pyrolysis : Products 1

    Distribution and Characterization 2 3

    Isah Yakub Mohammeda,f,*

    , Yousif Abdalla Abakrb, Mukhtar Musa

    c,f, Suzana Yusup

    d, Ajit Singh

    c 4

    and Feroz Kabir Kazie

    5

    6 aDepartment of Chemical and Environmental Engineering, the University of Nottingham 7

    Malaysia Campus, Jalan Broga, Semenyih 43500, Selangor Darul Eshan, Malaysia

    8

    9 bDepartment of Mechanical, Manufacturing and Material Engineering, the University of 10

    Nottingham Malaysia Campus, Jalan Broga, Semenyih 43500, Selangor Darul Eshan, Malaysia

    11

    12 cSchool of Biosciences, the University of Nottingham, Malaysia Campus, Jalan Broga, 43500 13

    Semenyih, Selangor, Malaysia 14

    15 dDepartment of Chemical Engineering, Universiti Teknologi PETRONAS, Bandar Seri Iskandar, 16

    31750, Tronoh, Malaysia 17

    18 eDepartment of Engineering and Mathematics, Sheffield Hallam University, City Campus, 19

    Howard Street, Sheffield S1 1WB, UK 20

    fCrops for the Future (CFF), the University of Nottingham Malaysia Campus, 21

    Jalan Broga, Semenyih 43500, Selangor Darul Eshan, Malaysia 22

    23

    Abstract 24

    This study provides first report on thermochemical conversion of residue from one of the 25

    underutilized crops, Bambara groundnut. Shells from two Bambara groundnut landraces KARO 26

    and EX-SOKOTO were used. Pyrolysis was conducted in a vertical fixed bed reactor at 500, 27

    550, 600 and 650 oC; 50

    oC/min heating rate and 5 L/min nitrogen flow rate. The report gives 28

    experimental results on characteristic of the feedstock, impact of temperature on the pyrolysis 29

    product distribution (bio-oil, bio-char and non-condensable gas). It evaluates the chemical and 30

    physicochemical properties of bio-oil, characteristics of bio-char and composition of the non-31

    condensable gas using standard analytical techniques. KARO shell produced more bio-oil and 32

    was maximum at 600 oC (37.21 wt%) compared to EX-SOKOTO with the highest bio-oil yield 33

    of 32.79 wt% under the same condition. Two-phase bio-oil (organic and aqueous) was collected 34

  • Page 2 of 24

    and analyzed. The organic phase from both feedstocks was made up of benzene derivatives 35

    which can be used as a precursor for quality biofuel production while the aqueous from KARO 36

    consisted sugars and other valuable chemicals compared to the aqueous phase from EX-37

    SOKOTO which comprised of acids, ketones, aldehydes and phenols. Characteristics of bio-char 38

    and composition of the non-condensable were also determined. The results show that bio-char is 39

    rich in carbon and some minerals which can be utilized either as a solid fuel or source of bio-40

    fertilizer. The non-condensable gas was made up of methane, hydrogen, carbon monoxide and 41

    carbon dioxide, which can be recycled to the reactor as a carrier gas. This study demonstrated 42

    recovery of high quality fuel precursor and other valuable materials from Bambara groundnut 43

    shell. 44

    Keywords: Bambara groundnut shell; Pyrolysis; Bio-oil; Bio-char; Non-condensable gas; 45

    Characterization 46

    *Corresponding author: Isah Yakub Mohammed ([email protected]) 47

    Abbreviations AAK Acid, aldehyde, ketones

    AMH Atomic mass of hydrogen

    AMO Atomic mass of oxygen

    C Carbon

    CFF Crops for the future

    DTG Derivative thermogravimetric

    EDX Energy dispersive x-ray

    ES Esters

    FTIR Fourier transform infrared

    GC Gas chromatograph

    GC-MS Gas chromatograph-mass spectrometer

    H Hydrogen

    HC Hydrocarbon

    HHV Higher heating value

    KBr Potassium bromide

    LHV Lower heating value

    LHw Latent heat of vaporization of water

  • Page 3 of 24

    N Nitrogen

    O Oxygen

    PH Phenolics

    RKO Karo Bambara shell

    S Sulfur

    SGR Sugars

    SKT EX-SOKOTO Bambara shell

    TCD Thermal conductivity detector

    TG Thermogravimetric

    TGA Thermogravimetric analyzer

    VAC Value added chemicals

    XRD X-ray diffraction 48

    49

    1. Introduction 50

    Bambara groundnut (Vigna subterranea (L.) Verdc), is an underutilized pulse legume cultivated 51

    primarily for its seeds, which are consumed in various forms of dishes, rich in protein (16-25 %), 52

    carbohydrate (42-60 %), lipid (5-6 %) (Brough and Azam-Ali, 1992), fibre (4.8 %) and ash (3.4 53

    %) (Brink et al., 2006). The fresh pods containing the seeds are boiled and eaten as snacks 54

    whereas the haulm, is used as livestock feed in most of the drier parts of Africa (Heller et al., 55

    1997). Legume such as Bambara groundnut with the ability to meet greater % of their N 56

    requirement as a result of atmospheric N fixation have been identified as potential alternative 57

    crops for mitigating climate change (Jensen et al., 2012) and offer an important alternative to the 58

    use of inorganic N fertilizers that are costly and unavailable to resource-poor farmers. The crop 59

    is predominantly grown by subsistence farmers over much of semi-arid Africa and is reported to 60

    spread into new production areas such as Indonesia, Malaysia and Thailand in the Asian 61

    continent and is increasingly becoming popular due to its ability to thrive under conditions which 62

    are too poor for other tropical legumes such as cowpea and groundnut (Mohale et al., 2014). 63

    Bambara groundnut shell constitutes the weight of the pod wall after the grain is removed and is 64

    highly variable depending on the pod yield and shelling percentage and sometimes could reach 65

  • Page 4 of 24

    up to 1000 kg ha-1

    . Musa et al. (2016) reported 374-896 kg ha-1

    of shell weight which 66

    corresponds to about 28.6-34.2 % of the total pod weight. Kishinevsky et al. (1996) reported a 67

    shelling percentage of 59-76 %, which translates to about 648-1107 kg ha-1

    ( 24-41 % of the pod 68

    weight) considering the average pod yield of 2700 kg ha-1

    reported by Collinson et al. (1999). 69

    Mwale et al. (2007) reported pod yield of 2980 kg ha-1

    and a shelling percentage of 80 %, which 70

    corresponds to a shell weight of 596 kg ha-1

    (20 % of the total pod weight). Despite the high 71

    weight of the crop’s shells, information on the utilization of this potential is limited and is 72

    seldom reported as a yield component due to lack of a specific value (Mwale et al., 2007). There 73

    is need to explore alternative ways to promote the utilization of this product which could address 74

    the problem of crop waste disposal and promote the integration of the crop into new production 75

    areas. Crop residue disposal in the sub-humid tropical regions such as Malaysia and Indonesia 76

    has been attracting a major global concern due to inappropriate methods adopted by the resource 77

    poor farmers, particularly in Indonesia, where farmers resort to the use of open burning as a 78

    cheap way of clearing crop residues from the farm and thereby causing air pollution (haze) to 79

    the neighboring countries and the entire region (Varkkey, 2013; Forsyth, 2014). 80

    Pyrolysis is currently one of the most promising thermochemical processes for converting 81

    biomass materials to products with high energy potentials. Typical pyrolysis products are bio-oil, 82

    bio-char and non-condensable gas (Alburquerque et al., 2016; Gómez et al., 2016). Distribution 83

    of pyrolysis products depends heavily on the process parameters such as pyrolysis temperature, 84

    heating rate and vapor residence time in the reactor. There are different types of pyrolysis 85

    namely; slow, intermediate and fast pyrolysis. Slow pyrolysis is also referred to as 86

    carbonization. It is carried out at a temperature up to 400 °C, solid residence time of 1 h to 87

    several hours, with product distribution of about 35 % bio-char, 30% bio-oil, and 35% non-88

  • Page 5 of 24

    condensable gas. Fast pyrolysis can produce up to 80 % bio-oil, 12 % bio-char, and 13 % non-89

    condensable gas at temperature around 500 °C, high heating rate, a short vapor residence time of 90

    about 1 s, and rapid cooling of volatile (Bridgwater, 2012; Mohammed et al., 2015a). For 91

    intermediate pyrolysis, 500-650 oC pyrolysis temperature and vapor residence time of 92

    approximately 10 to 30 s are typical operating conditions. About 40-60% of the total product 93

    yield is usually bio-oil, 15-25% bio-char and 20-30% non-condensable gas. Intermediate 94

    pyrolysis produce bio-oil with less reactive tar which can be used directly as fuel in engines and 95

    boilers, and dry char suitable for both agricultural and energy applications (Tripathi et al., 2016). 96

    To the best of our knowledge, thermochemical conversion of Bambara groundnut shells has not 97

    been reported in the literature. In this study, intermediate pyrolysis of Bambara shell was 98

    investigated in a fixed bed reactor and pyrolysis products were characterized. 99

    2. Materials and methods 100

    Shells from two Bambara groundnut landraces, Karo (RKO) and EX-Sokoto (SKT) were 101

    collected from Crops for the Future (CFF) field research Centre Semenyih, Selangor, Malaysia. 102

    The material was dried at 105 °C in electrical oven. The dried materials were then shredded in a 103

    Retsch® rotor beater mill (SM100, Retsch GmbH, Germany) to particle sizes between 0.2 mm 104

    and 2.5 mm and stored in airtight plastic bags for further analysis. Volatile matter and ash 105

    content on dry basis were determined. Fixed carbon was computed from the remaining bone dry 106

    sample mass. The biomass ash elemental composition was analyzed with energy dispersive X-107

    ray EDX (X-Max 20 mm2,

    Oxford Instruments, UK). Higher heating value (HHV) was 108

    determined using an oxygen bomb calorimeter (Parr 6100, Parr Instruments, Molin, USA). 109

    Elemental compositions were determined using CHNS/O analyzer (2400 Series II CHNS/O 110

    analyzer, Perkin Elmer Sdn Bhd, Selangor, Malaysia). The nature of chemical functional groups 111

  • Page 6 of 24

    were evaluated with Fourier transform infrared spectroscopy (Spectrum RXI, PerkinElmer, 112

    Selangor, Malaysia) using the potassium bromide (KBr) method. Potassium bromide (KBr) disc 113

    (13 mm diameter translucent) was made from a homogenized 2 mg sample in 100 mg KBr using 114

    a bench press (Carver 43500, Carver Inc.,USA) by applying 5.5 tonnes load for 5 min. Spectra 115

    were recorded with Spectrum V5.3.1 software within wave number range of 400 to 4000 cm-1

    at 116

    32 scans and a resolution of 4 cm-1

    . Thermogravimetric study was carried out using 117

    thermogravimetric simultaneous thermal analyzer (TGA) (STA 6000, Perkin Elmer Sdn Bhd, 118

    Selangor, Malaysia) in nitrogen atmosphere, flow rate 20 mL/min at temperatures between 30-119

    950 oC and heating rate of 20

    oC/min. About 10 mg (particle size of 0.2 mm) of sample was used. 120

    Intermediate pyrolysis study was conducted at 500, 550, 600 and 650 oC in a vertical fixed bed 121

    pyrolysis system as shown in Figure 1. The system consists of a fixed bed reactor made of 122

    stainless steel (115 cm long, 6 cm inner diameter), a distribution plate with 1.5 mm hole diameter 123

    which sit at 25 cm from the bottom of the tube, two nitrogen preheating sections, a cyclone, oil 124

    collector, gas scrubbers and water chiller operating at 3 oC attached to a coil condenser. 200 g of 125

    shell (bone dried, 2.5mm particle size) was placed on a gas distribution plate inside the reactor 126

    tube. The setup was heated in a vertical furnance at 50 oC/min under 5 L/min nitrogen flow. The 127

    reaction temperature was monitored with a thermocouple (K-type, NTT Heating, Sdn Bhd, 128

    Selangor, Malaysia). The reaction time was kept at 60 min after the temperature attained the set-129

    value. Pyrolysis vapor was condensed by passing through a condenser attached to the water 130

    chiller and condensate (bio-oil) was collected in a container. Bio-oil, bio-char and non-131

    condensable gas yields were calculated using Eq (1), (2) and (3). The experiment was repeated in 132

    triplicates and stanadard deviations were computed. 133

  • Page 7 of 24

    (1) -------- 100feed biomass ofweight

    collected oil-bio theofweight

    oil-bio(wt%) Yield

    134

    (2) ------ 100 feed biomass ofweight

    collectedchar -bio theofweight

    char-bio(wt%) Yield

    135

    (3) ----(2) Eq Eq(1)100gas econdensabl-non

    (wt%) Yield 136

    137

    Characterization of bio-oil, bio-char and non-condensable gas were carried out accordingly using 138

    analytical instruments as summerized in Table 1. FTIR of bio-oil was determined using a pair of 139

    circular demountable KBr cell windows (25 mm diameter and 4 mm thickness). Detail chemical 140

    composition of the bio-oil was determined using a gas chromatograph-mass spectrometer (GC-141

    MS) system equipped with a quadruple detector and column (30m x 0.25mm x 0.25µm). The 142

    oven was programmed at an initial temperature of 40 °C, ramp at 5 °C /min to 280 °C and held 143

    there for 20 min. The injection temperature, volume, and split ratio were 250 °C, 1 µL, and 50:1 144

    respectively. Helium was used as carrier gas at a flow rate of 1 mL/min. The peaks of the 145

    chromatogram were identified by comparing with standard spectra of compounds in the National 146

    Institute of Standards and Technology library (NIST, Gaithersburg, USA). 147

    The composition of non-condensable pyrolysis product was monitored offline. The gas sample 148

    was collected in a gas sample bag (Tedlar, SKC Inc., USA) and its composition analyzed using a 149

    gas chromatography equipped with stainless steel column (Porapak R 80/100) and thermal 150

    conductivity detector (TCD). Helium was used as a carrier gas and the GC was programed at 60 151

    oC, 80

    oC and 200

    oC for oven, injector and TCD temperature respectively. Bio-char ultimate 152

    analysis and inorganic composition were carried out following the procedure outlined for the 153

    feedstock characterization. Crystallographic structure in the produced bio-char was examined 154

    between 2Ө angle of 10°–70° at 25 mA, 45 kV, step size of 0.025°and 1.0 s scan rate. Surface 155

  • Page 8 of 24

    and structural characteristics of the bio-char were also determined . 156

    3. Results and Discussion 157

    3.1 Characterization of feedstock 158

    Characteristic of the feedstock is summarized in Table 2. Moisture content of SKT and RKO (as 159

    received) was 4.67 and 7.12 wt%. Higher volatile matter (76.24 wt%) and lower ash content 160

    (8.98 wt%) was recorded for the RKO compared to 73.83 wt% volatile matter and 10.10 wt% 161

    ash content in the SKT. Similarly, the ultimate analysis revealed that RKO has carbon and 162

    oxygen content of 37.36 wt% and 49.11 wt%, while 34.63 wt% carbon and 51.93 wt% oxygen 163

    was recorded for the SKT. RKO has higher heating value (HHV) of 20.46 MJ/kg relative to 164

    19.19 MJ/kg recorded for SKT. The higher energy content recorded for the RKO could be 165

    attributed to its lower ash and oxygen contents. These physiocchemical characteristics are similar 166

    to properties of peanut shell reported by Zhang et al. (2011). The authors recorded high volatile 167

    matter and heating value of 71.5 wt% and 18.1 MJ/kg respectively. Their ultimate analysis result 168

    showed higher carbon (47 wt%), and lower oxygen (34 wt%) and hydrogen (6 wt%) contents 169

    compared to that of the Bamabara shells. EDX analysis of the ash showed that the inorganic 170

    element in the RKO and SKT is in the following order potassium (K)>silica (Si)>Aluminum 171

    (Al)>phosphorus (P)>Chlorine (Cl)>iron (Fe)>calcium (Ca)>magnesium (Mg) and 172

    K>Cl>Si>Al>P>Mg respectively. The proportion of metallic elements in the RKO is similar to 173

    that of peanut reported by Kuprianov and Arromdee (2013). The authors also reported 174

    distribution of minerals in tamarin shell which is comparable to that of SKT except for the Si, 175

    which is higher in SKT. The result of thermogravimetric analysis (TGA) analysis is shown in 176

    Figure 2. The thermogravimetric (TG) and derivative thermogravimetric (DTG) profiles of both 177

    samples exhibited similar decomposition characteristics. The residual weight at the end of the 178

  • Page 9 of 24

    decomposition process was found to be 23 and 29 wt% for RKO and SKT respectively. From the 179

    DTG curves, two main peaks from ambient to 100 oC and from 240 to 420

    oC, which is attributed 180

    to evaporation of moisture and decomposition of holocellulose (hemicellulose and cellulose) 181

    component of the biomass. Hemicellulose which normally decomposes around 250 oC in a 182

    lignocellulosic material was not observed in this study. It means that Bambara shells have less or 183

    infinitesimal amount of hemicellulose. A noticeable peak around 200 oC was also observed, 184

    which is ascribed to the decomposition of extractives. The pattern observed at the temperature 185

    above 420 oC is attributed to the decomposition of lignin.This region exhibited unsteady thermal 186

    degradation, which is ascribed to the presence of different oxygen functional groups in the lignin 187

    (Mohammed et al., 2015b). TGA analysis of almond shell by Grioui et al. (2014) showed a 188

    residual weight fraction of 27.5 wt%. The DTG curve exhibited main peak at temperature 189

    between 220 and 400 oC, which was attributed to the decomposition of cellulose and 190

    hemicellulose. The decomposition temperature values observed in this study are aslo similar to 191

    those of other lignocellulosic biomass in the literature (Mohammed et al., 2015c). 192

    3.2 Pyrolysis product distribution and physicochemical properties of the bio-oil 193

    Product distribution from the pyrolysis of RKO and SKT at temperature between 500 and 650 194

    oC, 5 L/min nitrogen flow and heating rate 50

    oC/min is shown in Fig. 3. From Figure 3(a), total 195

    bio-oil yield from RKO increased from 33.26 to 37.21 wt% as the temperature increased from 196

    500 to 600 oC. This observation could be attributed to the decomposition of more lignin at such a 197

    high temperature. At 650 oC, the oil yield dropped to 27.69 wt%, which is due to secondary 198

    reactions of pyrolysis vapor at the elevated temperature (Mohammed et al., 2015a). The bio-oil 199

    collected consist of organic phase and an aqueous phase. The yield of organic phase increased 200

    from 9.8 wt% to 15.0 wt% at the temperature between 500 and 600 oC and latter declined to 9.3 201

  • Page 10 of 24

    wt% at 650 oC. On the other hand, a continuous decrease in the aqueous phase from 24.3 to 18.4 202

    wt% was recorded. Generally, the aqueous phase bio-oil is a product of holocellulose 203

    decomposition whose yield and composition is greatly affected by the inorganic minerals in the 204

    source biomass. The significant reduction in the bio-oil aqueous phase and increased non-205

    condensable yield observed could be attributed to the synergistic effect of the most abundant 206

    minerals in the biomass ash, which are Al and K (Table 2) since Si is an inert material and may 207

    not possess any catalytic activity (Mohammed et al., 2016a). The yield of non-condensable gas 208

    was 32.43 and 32.57 wt% at 500 and 550 oC but increased rapidity to 33.28 and 43.11 wt% at 209

    600 and 650 oC. The corresponding bio-char yield was 32.96, 32.26, 29.51 and 29.20 wt%. The 210

    high yield of non-condensable gas observed above could be mainly due to the secondary 211

    reactions of pyrolysis vapor at the elevated temperature (Mohammed et al., 2015a) since there 212

    was no substantial reduction in the biochar yield. Figure 3b shows pyrolysis product distribution 213

    from the SKT. The total bio-oil yield was 26.17, 32.18, 32.78 and 27.58 wt% at 500, 550, 600 214

    and 650 oC respectively. Similar to RKO, two phase bio-oil was collected from the pyrolysis of 215

    SKT and the corresponding organic and aqueous phases recorded were 5.90, 10.84, 11.17, 5.77 216

    wt% and 20.27, 21.34, 21.61, 21.81 wt%. The bio-char yield decreased from 39.91 wt% at 500 217

    oC to 28.56 wt% at 650

    oC while the corresponding yield of non-condensable gas increased from 218

    33.92 to 43.86 wt%. The high yield of non-condensable gas could be a resultant effect of a 219

    secondary reaction of pyrolysis vapor and devolatilization of the bio-char at high temperature. 220

    Comparing the product distribution between RKO and SKT, the highest total bio-oil yield of 221

    37.21 wt% was recorded at 600 oC from RKO compared to 32.78 wt% from SKT. A substantial 222

    decrease in the biochar yield (11.35 %) with increasing temperature (500-650 oC) was observed 223

    from SKT compared to 4.06 % recorded from the RKO. Higher non-condensable gas yield was 224

  • Page 11 of 24

    also recorded with SKT at all temperatures relative to that obtained from RKO. These 225

    observations can be attributed to the differences in the characteristics of the 226

    feedstocks(Mohammed et al., 2016b), as summerized in Table 2. Comparing the pyrolysis 227

    product distribution with products yield from other shells, Zhang et al. (2011) reported a total 228

    pyrolysis oil of 60 wt% from peanut shell at 500 oC with corresponding 19 wt% char and 15 229

    wt% non-condensable gas. Grioui et al. (2014) reported pyrolysis products distribution of 40 230

    wt% bio-oil , 47 wt% gas and 13 wt% solid. Pyrolysis oil yield from other biomass species have 231

    been reported by Lim et al. (2015) and (2016). They recorded bio-oil yield of 41.91-57.21 wt% 232

    and 37.98-43.66 wt% from Napier grass and sago waste respectively at 600 oC pyrolysis 233

    temperature. The variation in the products distribution can be linked to physicochemical and 234

    structural characteristics of the respective feedstock. Physicochemical properties of Bamabara 235

    shell bio-oil are summarized in Table 3. The pH value of organic phase from RKO and SKT was 236

    4.4 and 3.3, which indicate that the oils are less acidic relative to the almond bio-oil (pH = 3.1) 237

    reported by Grioui et al. (2014) (Table 3). Similarly, density of the bio-oils are relatively lower 238

    than the values reported in the literature for almond and peanut bio-oil. Ultimate analysis result 239

    (dry basis) calculated from the wet basis results using Eq (4), (5), (6) (de Miguel Mercader et al., 240

    2010; Mohammed et al., 2016c) and (7) (Mohammed et al., 2015d) indicated that C, H, O and 241

    HHV of the organic phase bio-oil from both Bambara shells are comparable to the elemental 242

    composition of peanut and almond shell bio-oil. 243

    )4(

    100

    21

    wetC

    dryC

    OH

    244

  • Page 12 of 24

    )5(

    100

    21

    AMOAMH2

    AMH2

    2wetH

    dryH

    OH

    OH

    245

    )6(

    100

    21

    AMOAMH2

    AMO

    2wetO

    dryO

    OH

    OH

    246

    )7(

    100

    21

    100

    21wetLHV

    dryHHV

    OH

    LHOH

    W

    247

    C, H and O is carbon, hydrogen and oxygen content in wt %; ηH2O is the amount of water in the 248

    oil (wt %); AMH and AMO is atomic mass of hydrogen and oxygen; LHw is the latent heat of 249

    vaporization of water (2.2MJ/kg) 250

    251

    3.3 Fourier-Transform Infra-Red (FTIR) 252

    FTIR spectra of chemical species in the bio-oil are shown in Figure 4. The common broad peak 253

    around 3439 cm-1

    implies that the samples contain chemical compounds with hydroxyl group (254

    HO ) such as water, alcohols and phenol (Bordoloi et al., 2015). The peaks between 2838 and 255

    2948 cm-1

    common to the organic phase bio-oils (Figure 4a and 4b) is HC stretching 256

    vibration, which indicate the presence of saturated hydrocarbon in the organic phase while the 257

  • Page 13 of 24

    peak between 2041 and 2132 cm-1

    common to all the samples is ascribed to the NC 258

    functional group (Guo et al., 2015). Vibrations observed between 1625 cm-1

    and 1707 cm-1

    in all 259

    the oil phases are attributed to OC , which signifies the presence of aldehydes, ketones or 260

    carboxylic acids. The vibration at a frequency of 1515 cm-1

    in both organic phases is ascribed to 261

    CC bond while the band around 1454 cm-1

    in all the samples is ascribed to HC bending, 262

    suggesting the presence of alkenes/aromatic hydrocarbons. The vibrations observed at 1364, 263

    1230 and 1273 cm-1

    signify NC stretching (Yorgun and Yıldız, 2015). The sharp band 264

    around 1112, 1093, 1050 and 1025 cm-1

    are due to OC vibration indicating the presence of 265

    alcohol and esters. The fingerprint at 755 cm-1

    is ascribed to aromatic HC bending vibrations 266

    while the fingerprint between 655 and 531 cm-1

    is ascribed to alkyl halide (Yorgun and Yıldız, 267

    2015). 268

    3.4 GC-MS analysis of the bio-oil samples 269

    Identification of detail chemical compounds in the bio-oil samples collected at 600oC was carried 270

    out by GC-MS. The chromatograms (Figure 5) show twenty (20) most abundant compounds in 271

    the bio-oils. The analysis revealed that the organic phase from both RKO and SKT consist of 272

    similar compounds such as alkenes, benzene derivatives, nitrogenated compounds, esters, 273

    aldehydes, ketones and acids in different proportions. A trace of sugar (L-Galactose) was also 274

    detected in the oil from RKO while none was observed in the oil from SKT. The aqueous phase 275

    from RKO consists predominant sugars and fine chemicals while that from SKT consist mainly 276

    acids, ketones, aldehyde and phenolics. In order to elucidate the variation of these compounds in 277

    the oil samples, the compounds were further grouped to hydrocarbon(HC), esters (ES), phenolics 278

    (PH), sugars (SGR), value added chemicals (VAC) and acid, aldehyde, ketones (AAK). From 279

    Figure 6, based on the GC-MS peak area (%), 7.09 % HC and 52.17 % PH was recorded in the 280

  • Page 14 of 24

    organic phase RKO while 66.06 %PH was observed in the SKT with 2.31 % HC. The high 281

    content of SGR (30.69 %) and low content of AAK (7.04 %) in the aqueous phase from the RKO 282

    compared to low SGR (4.42 % ) and high AAK (36.95 %) in the SKT is attributed to the 283

    catalytic effect of the respective biomass ash. Studies have shown that alkaline metals tend to 284

    catalyze pyrolysis reactions which usually lead to degradation and polymerization of pyrolysis 285

    product intermediate (Deshmukh et al., 2015; Bordoloi et al., 2015). Potassium (K) has been 286

    identified to have impacts on both the pyrolysis product distribution and the composition of bio-287

    oil. Studies by Shimada et al. (2008); Nowakowski et al. (2008); Mohammed et al. (2016a) have 288

    demonstrated that potassium strongly promoted the formation of low molecular weight 289

    compounds and inhibited the formation of sugar during biomass pyrolysis. Similar observations 290

    have also been reported by Wang et al. (2010) and Trendewicz et al. (2015) during the pyrolysis 291

    of pine wood and cellulose respectively. They also reported greater phenol yield with K as a 292

    catalyst during the pyrolysis. Therefore, the high proportion of PH, AAK and lower content of 293

    SGR in the oil from the SKT relative to RKO is ascribed to the higher K content in the SKT 294

    (72.57 wt% ) relative to 44.96 wt% in the RKO (Table 2). Compositions of bio-oil from almond 295

    shell bio-oil were mainly HC (alkanes and oleifins) as reported by Grioui et al. (2014) while 296

    peanut shell bio-oil constitutes about 59 % PH and 31 % AAK (Zhang et al., 2011). Lim et al. 297

    (2016) reported that bio-oil from Napier grass and sago waste consititute mainly PH and AAK 298

    and PH, AAK and trace of HC respectively. The authors linked the difference in the chemical 299

    composition to the presence of different mineral elements in the respective biomass. 300

    3.5 GC analysis of the non-condensable gas sample 301

    Composition the non-condensable gas collected at 600 oC is summarized in Table 4. The gas 302

    constitutes a high percentage of methane (CH4), hydrogen (H2), carbon monoxide (CO) and 303

  • Page 15 of 24

    carbon dioxide (CO2). The high composition of CH4 signifies thermal cracking mechanisms that 304

    produce small organic molecules during the pyrolysis. The proportion of CH4 recorded in the 305

    gas from RKO (9.43 vol%) was higher than that obtained from the SKT (7.42 vol%). This 306

    suggests that there is interaction between the minerals Al (19.61 wt%) and K (44.96 wt%) since 307

    Si is an inert material, (Table 2) in the biomass ash promoted the conversion of CH4 precursors 308

    to form hydrocarbons. This observation is in good agreement with the higher percentage of total 309

    hydrocarbons recorded in the bio-oil from RKO (18.28 % in both phases) relative to SKT (2.31 310

    %). On the other hand, higher composition of H2, CO and CO2 in the gas from SKT was recorded 311

    compared to RKO and could be linked to the higher percentage of K in the SKT ash. Study by 312

    Wang et al. (2010) has shown that K promotes production of H2, CO and CO2 during pyrolysis 313

    of lignocellulosic biomass. 314

    3.5 Characteristics of produced bio-char 315

    The properties of bio-char collected at 600 oC (Table 5) shows that both RKO and SKT produced 316

    bio-char rich in carbon. The atomic ratio, H/C and O/C were 0.03 and 0.14 in RKO-char while 317

    0.01 and 0.13 were recorded in SKT-char. The relative lower ratios observed in the SKT-char 318

    indicates that SKT produced higher carbonized char which connotes a greater degree of 319

    aromaticity. Comparing this characteristic with the individual biomass feedstock, H/C, O/C of 320

    RKO and SKT biomass was 0.3, 1.3 and 0.33, 1.5 (Table 2). The decrease in the ratio, 321

    particularly, the O/C in the produced bio-char was due to dehydration reaction during the 322

    pyrolysis (Al-Wabel et al., 2013). 323

    FTIR of the produced bio-char was compared with the source biomass as shown in Figure 7. The 324

    averaged spectra showed similar characteristic peaks for both RKO and SKT biomass. Similarly, 325

    common characteristic peaks were observed for the produced biochars. The peaks detected 326

  • Page 16 of 24

    between 3723 and 3505 cm-1

    in the feedstocks are assigned to different hydroxyl groups ( HO ), 327

    principally from the cellulose component of the biomass. The peak at 3334 cm-1

    is ascribed to 328

    HN functional group, indicating the presence of nitrogenous group in the feedstock while the 329

    peak observed at 3165, 3111 and 2960 cm-1

    are due to HC stretching vibrations of cellulose, 330

    hemicellulose and lignin in the biomass samples (Usman et al., 2015). These peaks disappeared 331

    in the produced bio-chars, confirming the decomposition of structural component of the biomass 332

    feedstock (Mimmo et al., 2014) while that observed around 3700 cm-1

    due to HO could be 333

    attributed to moisture adsorbed during sample preparations (Chen et al., 2016). The vibration at 334

    1658 and 1588 cm-1

    in the feedstock and produced bio-char is due to CC stretching while the 335

    band at around 1417 cm-1

    in all the samples is ascribed to HC bending. The band observed in 336

    the feedstock at 1246 and 1033 is ascribed to CC O and OC stretching vibrations of 337

    polysaccharides while the fingerprint at 545 and 462cm-1

    is due to alkyl halides (Mohammed et 338

    al., 2015b). These peaks completely disappeared in the produced bio-char confirming the 339

    degradation of polysaccharides and alkyl halides during pyrolysis. 340

    Physisorption analysis (Table 5) revealed that the produced bio-char is a porous material with a 341

    pore volume of 0.01 cm3/g (RKO-char), 0.2 cm

    3/g (SKT-char) with corresponding specific 342

    surface area (BET) of 1.2 and 1.72 m2/g. This observation is also evident in the SEM micrograph 343

    (Figure 8) which showed cracked, uneven and visible porous structure in the bio-char due to the 344

    high degree of devolatilization at high temperature (600 oC) during the pyrolysis. 345

    XRD analysis of the feedstocks and the produced bio-char is presented in Figure 9. The 346

    diffractogram showed main characteristic peak for both RKO and SKT biomass at 22.23o and 347

    22.09o respectively, which represent the crystalline component in the biomass, mainly cellulose 348

    (Timpano et al., 2015). This peak completely disappeared in the produced bio-char, 349

  • Page 17 of 24

    confirming complete decomposition of cellulose during the pyrolysis. New sets of peaks were 350

    identified in the produced bio-chars around 26, 28, 30, 31, 40 and 50o. XRD search and 351

    match reveal that the bio-char constituted predominantly sylvite mineral represented by the 352

    peaks in both chars around 28.63o, 40.65

    o and 50.43

    o (Usman et al., 2015). The peak observed 353

    around 26.83o and 30

    o, 31

    o indicate the presence of quartz and calcite (Yuan et al., 2011). 354

    Thermogravimetic analysis (Figure 10a) of the produced bio-char showed a total residual weight 355

    of 85.38 wt% and 86.49 wt% at 600 oC for SKT-bio-char and RKO-bio-char respectively 356

    compared to the individual feedstock with 37.67 wt% and 32.47 wt%. This result indicates that 357

    almost all the volatile matters in the biomass were completely volatilized during the pyrolysis. 358

    This observation is also confirmed in the DTG plot (Figure 10b) as all the peaks in the feedstock 359

    completely disappeared at 600 oC and above. The weight loss recorded at 600

    oC in the bio-char 360

    (13.51 wt%-RKO-char, 14.62 wt%-SKT-char) is due to combustion of the remaining organic 361

    material in the bio-char. At higher temperature (900 oC), the total weight loss was 78.77 wt% in 362

    RKO-char and 76.76 wt% in SKT-char, which correspond to about 7.72 wt% and 8.62 wt% 363

    weigth loss between 600 and 900 oC . This observation could be attributed to the decomposition 364

    of inorganic materials in the biochar at higher temperature (Azargohar et al., 2014). 365

    4. Conclusion 366

    Pyrolysis of Bambara groundnut shell was carried out in a fixed bed reactor. This study evaluates 367

    impact of temperature on the pyrolysis product distribution (bio-oil, bio-char and non-368

    condensable gas). It assesses the chemical and physicochemical properties of bio-oil, 369

    characteristics of bio-char and composition of the non-condensable gas. The bio-oil collected 370

    was a two-phase liquid, organic and aqueous phase. The organic phase from both shells have 371

    high energy content (25-29 MJ/kg) and could be used directly as a fuel in boilers or a precursor 372

  • Page 18 of 24

    for quality biofuel production. The aqueous phase particularly from RKO was predominantly 373

    made up of sugars and other value-added chemicals, which can be used as a raw material for 374

    production of consumer products and fine chemicals. Bio-char was a porous material rich in 375

    carbon, which may well be used as a solid fuel. The bio-char also consist of macronutrient and 376

    may be applied as a source of nutrient for plant growth. Non-condensable gas consists of 377

    methane, hydrogen, carbon monoxide and carbon dioxide which can be recycled to the reactor as 378

    an energy supplement for the endothermic pyrolysis reaction or as a carrier gas. Using Bambara 379

    shells as feedstock for bioenergy production will encourage more production of the groundnut, 380

    which will bring about food security, particularly in challenging environments where major 381

    crops are severely limited by adverse climatic conditions in addition to income generation. 382

    Acknowledgments 383

    The project was supported by the Crops for the Future (CFF) and University of Nottingham 384

    under the grant BioP1-005. 385

    386

    387

    388

    389

    390

    391

    392

    393

    394

    395

    396

    397

    398

    399

    400

    401

    402

    403

    404

    405

    406

  • Page 19 of 24

    407

    408

    409

    410

    411

    412

    413

    414

    415

    416

    417

    418

    419

    420

    421

    422

  • Page 20 of 24

    References 423

    Alburquerque, J. A., Sanchez, M. E., Mora, M., Barron, V., 2016. Slow pyrolysis of relevant 424

    biomasses in the Mediterranean basin. Part 2. Char characterization for carbon 425

    sequestration and agricultural uses. Journal of Cleaner Production 120, 191-197 426

    Al-Wabel, M.I., Al-Omran, A., El-Naggar, A.H., Nadeem, M., Usman, A.R.A., 2013. Pyrolysis 427

    temperature induced changes in characteristics and chemical composition of biochar 428

    produced from conocarpus wastes. Bioresource Technology 131, 374-379 429

    Azargohar, R., Nanda, S., Kozinski, J.A., Dalai, A. K., Sutarto, R., 2014. Effects of temperature 430

    on the physicochemical characteristics of fast pyrolysis bio-chars derived from Canadian 431

    waste biomass. Fuel 125, 90-100. 432

    Bordoloi, N., Narzari, R., Chutia, R.S., Bhaskar, T., Kataki, R., 2015. Pyrolysis of Mesua ferrea 433

    and Pongamia glabra seed cover: Characterization of bio-oil and its sub-fractions. 434

    Bioresource Technology 178, 83–89 435

    Bridgwater, A.V., 2012. Review of Fast Pyrolysis of biomass and product upgrading. Biomass 436

    and Bioenergy 38, 68-94. 437

    Brink, M., Ramolemana, G.M., Sibuga, K.P., 2006. Vigna subterranea (L.) Verdc. In: Brink, M. 438

    & Belay, G. (Editors). PROTA 1: Cereals and pulses/Cereales et legumes secs. [CD-Rom]. 439

    Wageningen, Netherlands 440

    Brough, S.H., Azam‐Ali, S.N., 1992. The effect of soil moisture on the proximate composition of 441

    Bambara groundnut (Vigna subterranea (L) Verdc). Journal of the Science of Food and 442

    Agriculture 60(2), 97-203. 443

    Chen, T., Liu, R., Scott, N.R., 2016. Characterization of energy carriers obtained from the 444

    pyrolysis of white ash, switchgrass and corn stover — Biochar, syngas and bio-oil. Fuel 445

    Processing Technology 142, 124-134 446

    Collinson, S.T., Azam-Ali, S.N., Chavula, K.M., Hodson, D.A., 1999. Growth, development and 447

    yield of Bambara groundnut (Vigna subterranea) in response to soil moisture. Journal of 448

    Agricultural science (Cambridge) 126, 307-318. 449

    Deshmukh, Y., Yadav, V., Nigam, N., Yadav, A., Khare, P., 2015. Quality of bio-oil by 450

    pyrolysis of distilled spent of Cymbopogon flexuosus. Journal of Analytical and Applied 451

    Pyrolysis 115, 43-50. 452

    de Miguel Mercader, F., Groeneveld, M.J., Kersten, S.R.A., Venderbosch, R.H., Hogendoorn, 453

    J.A., 2010. Pyrolysis oil upgrading by high pressure thermal treatment. Fuel 89, 2829–2837 454

  • Page 21 of 24

    Forsyth, T., 2014. Public concerns about transboundary haze: a comparison of Indonesia, 455

    Singapore, and Malaysia. Global Environmental Change 25, 76-86 456

    Gómez, N., Rosas, J.G., Cara, J., Martínez, O., Alburquerque, J. A., Sánchez, M.E., 2016. Slow 457

    pyrolysis of relevant biomasses in the Mediterranean basin. Part 1. Effect of temperature on 458

    process performance on a pilot scale. Journal of Cleaner Production 120, 181-190 459

    Grioui, N., Halouani, K., Agblevor, F.A., 2014. Bio-oil from pyrolysis of Tunisian almond shell: 460

    Comparative study and investigation of aging effect during long storage. Energy for 461

    Sustainable Development 21, 100-112 462

    Guo, Y., Song, W., Lu, J., Ma, Q., Xu, D., 2015. Wang S. Hydrothermal liquefaction of 463

    Cyanophyta: Evaluation of potential bio-crude oil production and component analysis. 464

    Algal Research 11, 242–247. 465

    Heller, J., Begemann, F., Mushonga, J., 1997. Bambara groundnut (Vigna subterranea (L.) 466

    Verdc.). Promoting the conservation and use of underutilized and neglected crops. 467

    Proceedings of the workshop on Conservation and Improvement of Bambara Groundnut 468

    (Vigna subterranea (L.) Verdc.), 14–16 November, Harare, Zimbabwe. 469

    Jensen, E.S., Peoples, M.B., Boddey, R.M., Gresshoff, P.M., Hauggaard-Nielsen, H., Alves, B.J., 470

    Morrison, M.J., 2012. Legumes for mitigation of climate change and the provision of 471

    feedstock for biofuels and biorefineries. A review. Agronomy for Sustainable Development 472

    32(2), 329-364. 473

    Kishinevsky, B.D., Zur, M., Friedman, Y., Meromi, G., Ben-Moshe, E., Nemas, C., 1996. 474

    Variation in nitrogen fixation and yield in landraces of Bambara groundnut (Vigna 475

    subterranea L.). Field Crops Research 48(1), 57-64. 476

    Kuprianov, V. I., Arromdee, P., 2013. Combustion of peanut and tamarind shells in a conical 477

    fluidized-bedcombustor: A comparative study. Bioresource Technology 140, 199–210 478

    Lim, C.H., Mohammed, I.Y., Abakr, Y.A., Kazi, F.K., Yusup, S. and Lam, H.L., 2016. Novel 479

    input-output prediction approach for biomass pyrolysis. Journal of Cleaner Production 30, 480

    1-11. 481

    Lim, C.H., Mohammed, I.Y., Abakr, Y.A., Kazi, F.K., Yusup, S. and Lam, H.L., 2015. Element 482

    characteristic tolerance for semi-batch fixed bed biomass pyrolysis. Chemical 483

    Engineering 45, 1285-1290 484

    Mimmo, T., Panzacchi, P., Baratieri, M., Davies, C. A., Tonon, G., 2014. Effect of pyrolysis 485

    temperature on Miscanthus (Miscanthus x giganteus) biochar physical, chemical and 486

    functional properties. Biomass and Bioenergy 62, 149-157. 487

  • Page 22 of 24

    Mohale, K.C., Belane, A.K., Dakora, F.D., 2014. Symbiotic N nutrition, C assimilation, and 488

    plant water use efficiency in Bambara groundnut (Vigna subterranea L. Verdc) grown in 489

    farmers’ fields in South Africa, measured using 15N and 13C natural abundance. Biology 490

    and Fertility of Soils 50, 307-319 491

    Mohammed, I.Y., Abakr, Y.A., Kazi, F. K., Yusuf, S., Alshareef, I., Soh, A.C., 2015a. Pyrolysis 492

    of Napier Grass in a Fixed Bed Reactor: Effect of Operating Conditions on Product Yields 493

    and Characteristics. BioResources 10 (4), 6457-6478. 494

    Mohammed, I.Y., Abakr, Y.A., Kazi, F.K., Yusup, S., Alshareef, I., Soh, A.C., 2015b 495

    Comprehensive characterization of Napier grass as a feedstock for thermochemical 496

    conversion. Energies 8(5), 3403-3417. 497

    Mohammed, I.Y., Abakr, Y.A., Kabir, F., Yusuf, S., 2015c. Effect of aqueous pretreatment on 498

    pyrolysis characteristics of Napier grass. Journal of Engineering Science and Technology 499

    10(11), 1487-1496. 500

    Mohammed, I.Y., Kazi, F.K., Abakr, Y.A., Yusuf, S., Razzaque, M.A., 2015d. Novel method for 501

    the determination of water content and higher heating value of pyrolysis 502

    oil. BioResources 10(2), 2681-2690. 503

    Mohammed, I.Y., Lim, C.H., Kazi, F.K., Yusup, S., Lam, H.L. and Abakr, Y.A., 2016a. Co-504

    pyrolysis of Rice Husk with Underutilized Biomass Species: A Sustainable Route for 505

    Production of Precursors for Fuels and Valuable Chemicals. Waste and Biomass 506

    Valorization, 1-11. 507

    Mohammed, I.Y., Abakr, Y.A., Kazi, F.K. and Yusuf, S., 2016b. Effects of pretreatments of 508

    Napier grass with deionized water, sulfuric acid and sodium hydroxide on pyrolysis oil 509

    characteristics. Waste and Biomass Valorization, 1-19. 510

    Mohammed, I.Y., Kazi, F.K., Yusup, S., Alaba, P.A., Sani, Y.M., Abakr, Y.A., 2016c. Catalytic 511

    Intermediate Pyrolysis of Napier Grass in a Fixed Bed Reactor with ZSM-5, HZSM-5 and 512

    Zinc-Exchanged Zeolite-A as the Catalyst. Energies 9 (4), 246. 513

    Murty, T.S., Scott, D., Baird, W., 2000. The 1997 El Nino, Indonesian forest fires and the 514

    Malaysian smoke problem: A deadly combination of natural and man-made hazard. Natural 515

    Hazards 21(2-3), 131-144. 516

    Musa, M., Massawe, F., Mayes, S., Alshareef, I., Singh, A., 2016. Nitrogen fixation and N-517

    balance studies on Bambara groundnut (Vigna subterranea L. Verdc) landraces grown on 518

    tropical acidic soils of Malaysia. Communications in Soil Science and Plant Analysis 519

    47(4), 533-542. 520

  • Page 23 of 24

    Mwale, S.S., Azam-Ali, S.N., Massawe, F.J., 2007. Growth and development of Bambara 521

    groundnut (Vigna subterranea) in response to soil moisture: 1. Dry matter and yield. 522

    European Journal of Agronomy 26(4), 345-353. 523

    Nowakowski, D.J., Jones, J.M., 2008. Uncatalysed and potassium-catalysed pyrolysis of the cell-524

    wall constituents of biomass and their model compounds. Journal of Analytical and 525

    Applied Pyrolysis 83, 12–25. 526

    Shimada, N., Kawamoto, H., Saka, S., 2008. Different action of alkali/alkaline earth metal 527

    chlorides on cellulose pyrolysis. Journal of Analytical and Applied Pyrolysis 81, 80–87. 528

    Timpano, H., Sibout, R., Devaux, M.-F., Alvarado, C., Looten, R., Pontoire, B., Martin, M., 529

    Legée, F., Cézard, L., Lapierre, C., 2015. Brachypodium cell wall mutant with enhanced 530

    saccharification potential despite increased lignin content. Bioenergy Research 8(1), 53-531

    67. 532

    Trendewicz, A., Evans, R., Dutta, A., Sykes, R., Carpenter, D., Braun, R., 2015. Evaluating the 533

    effect of potassium on cellulose pyrolysis reaction kinetics. Biomass and bioenergy 74, 15-534

    25 535

    Tripathi, M., Sahu, J.N., Ganesan, P., 2016. Effect of process parameters on production of 536

    biochar from biomass waste through pyrolysis: A review. Renewable and Sustainable 537

    Energy Reviews 55, 467–481 538

    Usman, A.R.A., Abduljabbar, A., Vithanage, M., Ok, Y.S., Ahmad, M., Ahmad, M., Elfaki, J., 539

    Abdulazeem, S.S., Al-Wabel, M.I., 2015. Biochar production from date palm waste: 540

    Charring temperature induced changes in composition and surface chemistry. Journal of 541

    Analytical and Applied Pyrolysis 115, 392-400 542

    Varkkey, H., 2013. Patronage politics, plantation fires and transboundary haze. Environmental 543

    Hazards 12(3-4), 200-217. 544

    Wang, Z., Wang, F., Cao, J., Wang, J., 2010. Pyrolysis of pine wood in a slowly heating fixed-545

    bed reactor: Potassium carbonate versus calcium hydroxide as a catalyst. Fuel Processing 546

    Technology 91, 942–950 547

    Yorgun, S., Yıldız, D., 2015. Slow pyrolysis of paulownia wood: Effects of pyrolysis parameters 548

    on product yields and bio-oil characterization. Journal of Analytical and Applied Pyrolysis 549

    114, 68-78. 550

    Yuan, J.-H., Xu, R.-K., and Zhang, H., 2011. The forms of alkalis in the biochar produced from 551

    crop residues at different temperatures. Bioresource Technology 102(3), 3488-3497. 552

  • Page 24 of 24

    Zhang, C., Zhang, R., Li, X., Li, Y., Shi, W., Ren, X., Xu, X., 2011. Bench-scale fluidized-bed 553

    fast pyrolysis of peanut shell for bio-oil production. Environmental Progress and 554

    Sustainable Energy 30, 11–18. 555

    556

    557